Facile backbone structure determination of human

Transcription

Facile backbone structure determination of human
Articles
Facile backbone structure determination of human
membrane proteins by NMR spectroscopy
npg
© 2012 Nature America, Inc. All rights reserved.
Christian Klammt1,2,6, Innokentiy Maslennikov1,2,6, Monika Bayrhuber1,3, Cédric Eichmann1,3, Navratna Vajpai1,3,
Ellis Jeremy Chua Chiu1,2, Katherine Y Blain1, Luis Esquivies1, June Hyun Jung Kwon2, Bartosz Balana4,
Ursula Pieper5, Andrej Sali5, Paul A Slesinger4, Witek Kwiatkowski1, Roland Riek1,3 & Senyon Choe1,2
Although nearly half of today’s major pharmaceutical drugs
target human integral membrane proteins (hIMPs), only
30 hIMP structures are currently available in the Protein
Data Bank, largely owing to inefficiencies in protein
production. Here we describe a strategy for the rapid structure
determination of hIMPs, using solution NMR spectroscopy
with systematically labeled proteins produced via cell-free
expression. We report new backbone structures of six hIMPs,
solved in only 18 months from 15 initial targets. Application
of our protocols to an additional 135 hIMPs with molecular
weight <30 kDa yielded 38 hIMPs suitable for structural
characterization by solution NMR spectroscopy without
additional optimization.
About 30% of the human protein-coding genes encode IMPs,
which have critical roles in metabolism, regulation, transport and
intercellular signaling. hIMPs are the targets of 50% of approved
therapeutic drugs; however, difficulties with the manipulation of
hIMPs have impeded the detailed functional and structural studies required to expedite drug development and discovery. These
difficulties are associated with hIMP expression, purification,
crystallization for X-ray structural studies, and isotopic labeling
and resonance assignment for solution NMR spectroscopy studies. Notably, cellular prokaryotic expression systems generally lack
compatible translocation machineries for hIMPs, and eukaryotic systems are expensive and difficult to handle. Consequently,
only 30 structures of hIMPs are currently deposited in the Protein
Data Bank (PDB).
Recently, Escherichia coli–derived cell-free expression systems
have proven effective in overcoming many limitations inherent to in vivo expression in prokaryotic hosts1. In the absence
of compartmentalization in a hydrophobic milieu, IMPs produced in a cell-free expression system form precipitates that can
be subsequently solubilized in mild detergents. We named this
mode of expression precipitating cell-free (P-CF) expression1.
Alternatively, inclusion of a detergent or a lipid can effect direct
expression of solubilized IMPs2–5. We have extensively optimized
P-CF expression for IMP production and had demonstrated efficient production of natively folded protein6. Other studies have
also shown cell-free expression of fully functional G protein–
coupled receptors and transporters7–11.
Transverse relaxation optimized spectroscopy (TROSY)-based
experiments have expanded the applicability of three-dimensional
(3D) structure determination by solution NMR spectroscopy to
large systems12, including micelle-bound membrane proteins13–17.
The tremendous adaptability of cell-free expression makes it ideally suited to the isotopic labeling strategies used for these experiments. In particular, the cell-free combinatorial dual-labeling
(CDL) strategy6 has greatly facilitated the usually laborious
sequential assignment of IMP resonances. Furthermore, technological limitations in the acquisition of the requisite long-range
distance constraints for 3D structure determination have been
overcome thanks to the measurements of paramagnetic relaxation
enhancement (PRE) caused by an exogenous or covalently bound
paramagnetic group18–21 and the measurements of long-range
nuclear Overhauser effects (NOE) for deuterated and selectively
protonated proteins22 solubilized in deuterated detergents. Here
we describe an application of a highly effective and fast strategy,
which combines NMR spectroscopy and cell-free expression, to
determine the structure of six hIMPs.
RESULTS
Expression screening and NMR spectroscopy
We surveyed the hIMP proteome for favorable targets for
solution NMR spectroscopy structural studies (Fig. 1a) and initially selected 15 moderately sized (<20 kDa), polytopic (two or
more membrane crossings) hIMPs (Fig. 1b–f). We expressed all
but one of these in our E. coli–derived P-CF system at high levels (>1 mg per 1 ml of reaction mixture) (Fig. 1c,d). The proteins were suitable for subsequent characterizations without
1Structural Biology Laboratory, Salk Institute for Biological Studies, La Jolla, California, USA. 2Joint Center for Biosciences at Gachon University of Medicine and Science,
Incheon, Korea. 3Laboratorium für Physikalische Chemie, ETH Zürich, Zürich, Switzerland. 4Clayton Foundation Laboratories for Peptide Biology, Salk Institute
for Biological Studies, La Jolla, California, USA. 5Department of Bioengineering and Therapeutic Sciences, Department of Pharmaceutical Chemistry and California
Institute for Quantitative Biosciences, University of California at San Francisco, San Francisco, California, USA. 6These authors contributed equally to this work.
Correspondence should be addressed to S.C. ([email protected]).
Received 20 January; accepted 25 April; published online 20 May 2012; doi:10.1038/nmeth.2033
834 | VOL.9 NO.8 | AUGUST 2012 | nature methods
G att2
1A
D
D
O
IG
H
*
Solution
structure
15
12
9
2
G
2
G
X
3
G
X
2
F
12
9
6
2
0
7
3
1
5
2
G
X
4
G
X
4
G
e 15
6
3
3
G
X
15
0
15
2
2
4
F
3
F
f
1
0
2
G
15
12
9
3
P
4
F
9
5
Poor
d
2
G
X
Fair
11
TMH: 2
NMR: F
STR:
Good
Accelerated
assignment
*
*
*
6
3
0
0
1
Figure 1 | Workflow for screening and evaluation of 15 selected hIMPs for structural studies. (a) hIMP genes (G) are cloned into suitable Gateway-adapted cellfree vectors (Online Methods). P-CF expression is screened using a 24-well preparative high-throughput (pHTP) setup, followed by verification of expression,
efficiency of detergent solubilization and suitability for NMR spectroscopy. Structure determination involves P-CF expression of uniformly or selectively labeled
sample, protein solubilization, assignment using CDL and standard approaches, and collection of PRE and NOE data for structure calculation. (b,c) Western
blot (b) and Coomassie stain gel (c) of indicated proteins (hIMPs with determined structures are labeled in green). TMH, numbers of predicted transmembrane
helices; NMR, NMR spectral quality information labeled as good (G), fair (F) or poor (P); structure, ‘X’ marks hIMPs with determined solution structure. Asterisks
indicate the position of expressed proteins. (d–f) Summary of evaluation for the 15 proteins. Expression of all 15 proteins was verified by western blot. Number of
proteins classified according to cell-free expression levels (d). Number of proteins solubilized with each of the tested detergents: 70 mM sodium dodecyl sulfate
(SDS), 42 mM 1-myristoyl-2-hydroxy-sn-glycero-3-[phospho-rac-(1-glycerol)] (LMPG), 100 mM n-decylphosphocholine (FC10), 100 mM n-dodecylphosphocholine
(FC12), 250 mM n-decyl-β-d-maltoside (DM), mixture of 196 mM n-dodecyl-β-d-maltoside (DDM) with 41 mM cholesteryl hemisuccinate (CHS) and 100 mM
lauryldimethylamine-N-oxide (LDAO) (e). Number of proteins classified according to NMR spectral quality (f).
additional purification. The lipid-derived detergent 1-myristoyl2-hydroxy-sn-glycero-3-[phospho-rac-(1-glycerol)] (LMPG)
was the most effective of the seven detergents screened for
solubilizing the protein precipitates produced by P-CF expression
(Fig. 1e). We measured [1H-15N]TROSY–heteronuclear single quantum coherence (HSQC) spectra from the solubilized, uniformly
15N-labeled hIMPs and evaluated their quality (good, fair or poor;
Fig. 1c,f) according to the number of visible glycine backbone and
tryptophan indole H-N resonances, the total number of cross-peaks,
the chemical shift dispersion and the uniformity of line shapes.
Characterization and NMR signal assignment of six hIMPs
From our initial 15 hIMP preparations, nine produced good
[1H-15N]TROSY-HSQC spectra and were suited for compre­hensive
NMR spectroscopy studies. For six of these nine hIMPs, we assigned
105
c
G32
G85
G42
G81
G7
G15
G40
G25
110
TMEM14A
G68
a
HIGD1A
125
G37
G74
G53
G33
G34
G78
T25
T5
S11
S83
S18
120
125
T3
S50
T54 T76
Q68
S57
S86
S18
F87
E86
M75
R85 K49 I21 R22 V32 V30
N53 E13
V65
Y43
V7 A39 W88
I34
K47
V77 V71
R64
L48 L20 R22 I40 R51
S8
A38
V41 R22
Y84
F70
L45
Y81
M56
V72
D4 K55
F29 I58 L60
A89
A35
K92
A66
A24
9.8 9.7 9.6
W88
128
T64
C71
S51
S54
S28
S2
V83
tag
S28 S74 S87
S16
R93
V46
L96 M67S47 I30
V35 I77 V95
S56
F61
F18
L98
F58
T13 M67
I91
M1 V12
M90
Y44
I4
L63
tag
L94
Y20 K76
F34 L11 R45
A64
L92
I66
L38
F71
V55 D49
V69
F14 M78
R51
K50
A9
L82
D52
D2
L99
R73
K72
I17
V26 L3
R70 L86
R23
C37 A10
V53 K21
A39
F62
L29 F6 A31 K75
L33 A43
L57
K54 A60
L29
A84
A80
T55
E27
I62
N4
R48 V66 V40
R50
Y44
T83
C16
V81V89
K24
V41
V31
Q69
M92
L21 V73 Q94
D15 R45
C38 R26
H63 I46
R91
L22 R23 D13 K20
Y85 A43
L30
K90
M77 E14
I76
A68 L49
L61
D87
L39
M57
Y82
E19 W7
Y47
H60 E98 R5 M84 I33 R52 A93
I59
A42
A70
R6 L35 Y88 A72
V17
A75
R65
A67
7.0 6.8
K56
A3
118 W8
10.3 10.2
A96
A80
L78
W7
W8
K99
8.8 8.6 8.4 8.2 8.0 7.8 7.6
1
105
122
G21
Figure 2 | NMR spectral quality and N-H backbone assignment for six hIMPs.
(a–f) [1H,15N]TROSY-HSQC spectra with assignment of NH cross-peaks for six
hIMPs (Online Methods). Insets, tryptophan indole NH cross-peaks (bottom
left), and high-field shifted backbone NH cross-peaks (bottom right).
G5b
G77
T37
TMEM141
S5
S32
S9 N58
S100
S80
A63
130
f
G94
G90
G49
G45
G25
G41
120
125
T104
W62
W90
T68
T4
E94
S28
K16
S75
H17 S83 C26 V66 C86
S103
T37
K55
R54
M32 V13
M50
Q53
Q97
I52
F46
N88
L91
Q49 Q27
V74
A15
V81
V68 K85
F31
D11
F36
A12 K33 F51 W90 Y58 V9 L65
V78
L93
Y23 R8
L6
N87
M44
A25
K100
D101
R8b R80
A67
A24
H29 E22
A14 V2 E82 Y76
F92
Q106
A70
L4
N3
R102
V73
A12
L48 R107
S108
D105
L60
E84
L64
F38
L98
L20
L4 V35
A30
10.1
9.9 W62
S7
G24
G84
G70
T79
129
S34
I75
H11
10.2 10.1
W62
8.8 8.6 8.4 8.2 8.0 7.8 7.6
1
S5
S6
S56
S83
S23
S96
S106
tag
N61
V30 S69
S46
V64
F108
M76
I92
T71
F13
V21
Y53 M99
I27 N109
W12 V100
L43 M78 L26 L20 V7 M107 tag
D3
A67
W63 Q57
K31
A18
A40 Q54
Y29
L66
A39 A73
L91
Y17
L55
L50 A93
L10
F65
L97
V35
F15
V62
I75 L72
A95
A32
A48
A52
R110 A22 A48
tag
A89
L42
L38
H112
R60
V105
W12
130
F44
G33
G14
G28
G104
G16
T40
S63
115
TMEM14C
G41
G74
G51
G34
G96
S52
T82
S73 T92
129
G5a
G43
G19
G88
G20
G6
tag
T53
T26
S38
S103
I56
M34
V14
M51
L94
S66
V17 I86 V68 S45 L98
Q72 L31
V24
V18
V29
tag
K49
V85
L28
V81
A77 M50
I70
F36
E4
V74A55 V62
R11
W97
V21 A12
A43
A69
A16
D8 L71
A15
A48
W7
A90
F89 V39
A13
A76
A64
L67 I41 K84 V27
A47
I46
L79
A57
S104 A96
A93
A30
tag
A54
A33
A42
10.2 10.1
A23
T95
G47
110
V9
8.6 8.4 8.2 8.0 7.8 7.6 7.4 7.2
H (p.p.m.)
e
G71
G99
tag
G59
S83
W97
120
130.5
tag
G35
G87
G61
W7
S58
130.0
L9
129
15
N (p.p.m.)
G6
115
Y8
L88
G97
G95
G2
T59
120
G36
G73 G80
110
G19
T65
HIGD1B
G52
G60
S44
G79
G69
G17
G40
G75
G25
G5
FAM14B
G78
G10
N (p.p.m.)
G36
G44
b
G65
G36
M89
105
d
G19
G24
115
15
npg
© 2012 Nature America, Inc. All rights reserved.
Structure
Determination
CDL
NMR
(iv)
Dialyzer with
LMPG
sequential
samples
hIMP precipitate
( H N C),
PRE, NOE
15 h in RMX (1–4 ml)
15 13
2 15
( N C), ( H N)
(d27-LMPG) (v)
data
15
( N), (CDL),
15
(selectively N)
2 15 13
*
*
*
*
DM
Detergent
solubilization
*
*
*
DDM+CHS
LDAO
LMPG
*
FC12
Preparative
scale
NMR
suitability
*
*
*
14
FC10
(iii)
16
SDS
15 h
11
LMPG
NMR
samples
15
( N)
TROSYHSQC
c
0.5 mg ml–1
detergent-
Reaction mixture solubilization
Isotopic labeling Feeding mixture
(0.75–1.4 ml) Dialysis (RMX) (50–100 µl)
screen
(uniform, CDL)
membrane
Expression level,
detergent
solubilization
1 mg ml–1
SDS gel
14
2 mg ml–1
(24 × 70 µl)
(ii)
Expression
verification
western
blot
>3 mg ml–1
Screening
Anti His6
(i)
MW (kDa)
24-well pHTP scale
Expression
screening
MW (kDa)
16
Number of hIMPs
Cloning
att1 G att2 Thr Xa Stagll His6
p23-GWN (NMR) for NMR screening
hIMPencoding
gene in
cell-free
vector
AP
b
p23-GWT (tagged) for verification
BL
C
a
LP
P1
H
IG
D
1
TM B
EM
PS 14A
EN
EN
FA
M
14
TM B
EM
1
TM 41
EM
1
TM 4C
EM
14
B
LE
PR
O
SC TL
1
D
5
AT
P5
G
1
C
N
IH
4
C
14
or
f1
Articles
129.0
129.5
H82
A93
8.6 8.4 8.2 8.0 7.8 7.6 7.4 7.2
H (p.p.m.)
nature methods | VOL.9 NO.8 | AUGUST 2012 | 835
npg
© 2012 Nature America, Inc. All rights reserved.
Articles
Figure 3 | Solution NMR spectroscopy
structures and long-distance constraints used
in structure calculations. (a) Stereo view of
the TMEM14A structures calculated using
separate sets of NOE- and PRE-based longrange distance constraints. The best structure
calculated using NOE constraints (red) was
superimposed with the best 20 structures
calculated using PRE-based constraints
(black). Long-range NOE constraints are shown
with blue dotted lines between heavy atoms
(instead of attached protons) for clarity of
the figure. The structures are superimposed
by backbone atoms of transmembrane
helical regions. The backbone of the region
Gly24–Leu98 is shown. Side chains with the
long-range NOE contacts are also shown for
NOE-based structure. (b) Ribbon representation
of solution NMR spectroscopy structures
of six hIMPs and long-distance constraints
used in structure calculation. The long-range
distance constraints obtained from PREs
measurements are indicated by straight lines
colored according to the color of the spinlabeled transmembrane helix (first is colored
in green, second in blue and third in orange).
Additional non-transmembrane helices and
PRE constraints obtained with the spin labels
located outside the transmembrane regions are
colored in gray. The structures are shown in
pairs with 90° vertical rotation.
a
resonances (Fig. 2) using the CDL strategy6 (Supplementary Table 1)
and conventional sequential assignment. The percentages of resonances assigned for backbone and side-chain atoms for the six
proteins are summarized in Supplementary Table 2. Static light scattering coupled with size-exclusion gel chromatography and refracting
index measurements revealed that these hIMPs were monomeric in
LMPG micelles (Supplementary Fig. 1). We verified that H1GD1B
was also monomeric in three different detergents: LMPG, dodecylphosphocholine (FC12) and n-dodecyl beta-maltoside (DDM)
(Supplementary Fig. 2). We also verified that proteins were micelleembedded and determined the regions of the target proteins embedded into the LMPG micelle by analyzing the PRE effect caused by the
hydrophilic spin-labeled probe (Supplementary Fig. 3).
NMR spectroscopy structure determination of hIMPs
We used the chemical shift index calculated for 13Cα, 13Cβ and
13CO atoms to localize the helical regions of the proteins. We calculated backbone structures of the six hIMPs based on long-range
distance constraints obtained from PRE measurements of the spinlabeled proteins using single cysteine mutants (Supplementary
Figs. 4 and 5 and Supplementary Table 3). To validate the PREbased structure calculations, we calculated the spatial structure
of TMEM14A using PRE and independently using long-range
distance constraints obtained from NOEs. The structures were
similar, with average r.m.s. deviation between the backbone atoms
of the transmembrane helical regions in the PRE- and NOE-based
structures equal to 3.05 Å (Fig. 3a). The NOE-based structure of
TMEM14A comprised a more tightly packed helical bundle than
the PRE-based structure, whereas the orientation and topology
of the transmembrane helices were very similar. The reason for
the less tight packing of the bundle in the PRE-based structure
836 | VOL.9 NO.8 | AUGUST 2012 | nature methods
b
HIGD1A
HIGD1B
TMEM14A
TMEM141
TMEM14C
FAM14B
lies in the impossibility of calculating short (<12 Å) PRE constraints using nitroxide paramagnetic spin label and in the lower
precision of this type of constraints as compared to NOE-based
constraints (Online Methods). To determine whether the helical orientation can be unambiguously determined from the PRE
data, we calculated the PRE constraints error for generic structures with transmembrane helices rotated around the helical axes.
The structures with the low cumulative error function (smaller
than 1 Å2) calculated for PRE constraints were located within
a 20° rotation range from the starting structure (angles 0, 0)
(Supplementary Fig. 6). PRE distances quality factor did not
exceed 13.6% among the calculated hIMP structures (Table 1),
which shows good agreement between PRE-derived distances
used in structure calculation and distances back-calculated from
the structures. NMR spectroscopy experimental data, including
structural and refinement statistics for the calculated structures,
are summarized in Table 1. The long-range distance constraints
used for structure calculation are illustrated in Figure 3b.
NMR spectroscopy structures of six hIMPs
All six hIMP structures were helical bundles with helix
lengths and exposed hydrophobic faces consistent with the
bilayer-embedded localization of these proteins (Fig. 4). In
agreement with the numbers of predicted transmembrane
helices (Fig. 1c), HIGD1A, HIGD1B and TMEM141 had two
transmembrane crossings, whereas TMEM14A had three transmembrane crossings. compared to the prediction of three and
four transmembrane helices, FAM14B and TMEM14C had two
and three transmembrane crossings, respectively (Fig. 4 and
Supplementary Fig. 7). The first transmembrane helix of both
FAM14B and TMEM141 was severely kinked (by about 50°;
Articles
Table 1 | Summary of NMR spectroscopy data and statistics for the calculated sets of 20 lowest-energy structures
npg
© 2012 Nature America, Inc. All rights reserved.
TMEM14A
Protein size (amino acids)
Transmembrane segments
(predicteda)
Global rotational correlation
timeb (ns)
Cysteine mutants
NMR constraints
NOE (long-range NOEd)
PRE: upper/lower
Number of PRE per restrained
residue: upper/total restrained
distances
Hydrogen bonds
Dihedral angle: Phi/Psi
Structure statisticse
Violations (mean ± s.d.)
Distance, sum (Å)
Distance, average (Å)
Maximum distance
violation (Å)
Dihedral angle, sum (°)
Maximum dihedral angle
violation (°)
PRE distances Q-factor (%)
Average backbone r.m.s.
deviation (Å)
Ramachandran plot
statisticse,f
Most favored regions (%)
Additionally/generously
allowed (%)
Disallowed (%)
Deviations from idealized
geometryg
Bond length (Å)
Bond angles (°)
Equivalent resolutionf,g (Å)
HIGD1A
HIGD1B
NOE
FAM14B
TMEM141
TMEM14C
93
2 (2)
99
2 (2)
99
3 (3)
PRE
104
2 (3)
108
2 (2)
112
3 (4)
25.1
28.8
25.5
21.6
20.7
25.5
6
6
7
8
5
9
40c
156/186
3.63/5.23
47c
224/240
4.07/6.34
651 (18d)
0
0
0
334/467
4.64/7.38
49c
195/389
3.98/8.39
153
162/235
1.80/2.61
55c
283/479
4.56/8.76
41
43/43
44
53/55
44
72/72
44
72/72
51
45/49
60
86/90
58
62/61
2.80 ± 0.10
0.19 ± 0.03
0.24 ± 0.01
1.35 ± 0.11
0.17 ± 0.04
0.25 ± 0.03
2.53 ± 0.39
0.15 ± 0.06
0.23 ± 0.03
2.42 ± 0.18
0.13 ± 0.06
0.21 ± 0.03
3.11 ± 0.16
0.26 ± 0.14
0.51 ± 0.02
1.15 ± 0.07
0.09 ± 0.04
0.14 ± 0.01
3.29 ± 0.36
0.21 ± 0.06
0.31 ± 0.04
1.36 ± 0.55
0.56 ± 0.13
1.02 ± 0.67
0.51 ± 0.30
3.78 ± 1.24
0.90 ± 0.29
2.11 ± 0.60
0.84 ± 0.32
4.87 ± 0.72
0.75 ± 0.19
3.73 ± 1.37
0.48 ± 0.10
4.78 ± 1.39
0.73 ± 0.21
12.3
1.52 ± 0.59
13.4
0.73 ± 0.19
NA
1.72 ± 0.65
12.1
1.82 ± 0.48
13.6
1.11 ± 0.42
NA
3.52 ± 1.51
13.1
1.42 ± 0.37
71.6
19.6/6.2
67.3
23.1/7.1
73.0
17.6/7.1
70.6
22.4/4.9
70.7
21.9/4.8
80.9
15.2/2.8
71.9
17.8/7.2
2.5
2.5
2.3
2.1
2.7
1.1
3.1
0.013
1.6
3.3
0.012
1.6
3.3
0.013
1.5
2.8
0.013
1.5
3.1
0.010
1.6
3.2
0.012
1.6
3.0
0.013
1.5
3.1
aNumber of predicted transmembrane helices is given in parentheses. bCalculated for total mass of protein-detergent complex using the Stokes-Einstein equation. cInter-residue sequential
(|i – j| = 1) constraints only. dNumber of long-range (|i – j| > 4) NOE constraints is given in parentheses. eCalculated for transmembrane bundles. fCalculated by Procheck program31.
gBased on Ramachandran plot quality assessment.
NA, not applicable.
Supplementary Table 4). For HIGD1A, HIGD1B, TMEM141
and TMEM14A, the first transmembrane helix was preceded by
an amphiphilic N-terminal helix; this helix is presumably located
at the micelle-water interface, but we could not define its exact
orientation relative to the transmembrane helical bundles. The
connecting loops between the transmembrane helices of FAM14B
and second and third transmembrane helices of TMEM14C contain an amphiphilic helix, which lies roughly perpendicular to the
preceding transmembrane helix. As a moderate PRE effect caused
by the soluble paramagnetic agent was detected for backbone
amides of the amphiphilic helices (Supplementary Fig. 3), we
can assume that the helices are located close to the surface of the
LMPG micelle. TMEM141 has remarkably elongated transmembrane helices of 34 and 33 amino acids. The N-terminal part of
the first transmembrane helix and C-terminal part of the second
one protruded from the micelle as confirmed by a moderate PRE
effect from the soluble paramagnetic agent (Supplementary Fig. 3d).
The three-helical bundles of TMEM14A and TMEM14C were
tightly packed with interhelical distances less than 8 Å, whereas
the two-helical bundles of HIGD1A, HIGD1B, FAM14B and
TMEM141 were more loosely packed with interhelical distances
exceeding 8 Å and relatively few interhelical van der Waals contacts localized close to the ends of the helices. The parameters
describing packing of the helices (pair-wise angles between the
transmembrane helices and bending angles of the transmembrane helices) are listed in Supplementary Table 4. The backbone
structures of HIGD1A, HIGD1B, FAM14B, TMEM141,
TMEM14A and TMEM14C reported here account for 20% of
hIMP entries currently in PDB.
The structures of these proteins could immediately serve
as a model for a substantial portion of the membrane proteome. A search of the UniProtKB database of all known
­protein sequences identified 609 unique protein sequences
with sequence identity greater than 30% to at least one of the
six hIMPs. Based on the determined structures, we calculated
structural models for these 609 unique protein sequences.
nature methods | VOL.9 NO.8 | AUGUST 2012 | 837
Articles
HIGD1A
HIGD1B
FAM14B
TMEM141
TMEM14C
npg
© 2012 Nature America, Inc. All rights reserved.
TMEM14A
Figure 4 | Solution NMR spectroscopy structures of six hIMPs. Structures
were calculated by Cyana using distance information obtained from NOE
and PRE measurements. The first transmembrane helix is in green, second
in blue and third in orange. Additional non-transmembrane helices are
in gray. Shown are ribbon structures (left), superposition of the best
20 backbone structures (middle) and their 90° vertical rotation (right).
Additional lower ­accuracy ­structural information is provided
for additional 380 protein sequences at sequence identities
below 30% (Supplementary Table 5).
hIMP-specific antibody generation by a P-CF product
We demonstrated that the P-CF–expressed hIMP elicit the production of highly specific polyclonal antibodies. A rabbit polyclonal
antibody (Eton Bioscience) was generated to P-CF–expressed
and detergent-solubilized HIGD1A. The anti-HIGD1A antibody
preparation recognized an overexpressed HIGD1A-GFP fusion
in HEK293T cells and endogenously expressed HIGD1A in both
838 | VOL.9 NO.8 | AUGUST 2012 | nature methods
HEK293T cells and hippocampal neurons but did not recognize the homologous (43% sequence identical) protein HIGD1B
(Supplementary Fig. 8).
Screening a broad range of hIMPs
The success of our preliminary studies spurred a more extensive coverage of the hIMP proteome. From a library of 3,270
hIMPs23, we selected an additional 135 targets in the 10–30 kDa
range for P-CF expression, solubilization screening and preliminary NMR spectroscopy analysis (Supplementary Fig. 9).
Overall, 111 (74%) of 150 targets expressed at considerably high
levels (>1 mg ml−1 of cell-free reaction mixture; Supplementary
Fig. 10 and Supplementary Table 6), and we used LMPG to solubilize all of them. From an analysis of [1H-15N]TROSY-HSQC
spectra, we found that 38 of 100 evaluated by NMR spectro­scopy
targets, including the six hIMPs with solved structure, were
adequate for structural studies without additional optimization
(Supplementary Figs. 11 and 12).
DISCUSSION
The described structural studies of hIMPs demonstrate the technological synergy between cell-free expression and CDL-aided
NMR spectroscopy analysis. It has been shown previously that
PRE-based constraints can be used for structure determination
of β-barrel membrane proteins19 and small single-crossing helical
membrane proteins18. They also have been used as additional
long-range distance constraints for structure determination
of multihelical membrane proteins14,24,25. Our results demonstrate that PRE-derived distances can be used as a single source
of long-range constraints for structure determination of multihelical membrane proteins. A careful design of the mutants for
spin labeling and a meticulous analysis of PRE data, however, are
necessary because changes in protein conformation induced by
mutation, high mobility of the spin-labeled residue and sparse net
of long-range PRE data may affect accuracy of the structures.
The biological functions of the six hIMPs we characterized here
have not yet been fully defined, and we do not know whether their
structures are in the functionally relevant state. Nevertheless,
knowledge of the backbone scaffolds of the proteins may provide structural insights for, for example, site-specific mutagenesis, which would help understand the fundamental functional
roles of these proteins. HIGD1A and HIGD1B are likely associated with response to hypoxia26. HIGD1A has been found to be
upregulated in hypoxia27; HIGD1B in prolactinomas is speculated to be associated with increased tumor hypoxia tolerance,
angiogenesis and drug resistance28. FAM14B, a member of
the FAM14 family encoded by interferon-stimulated gene 12c
is localized in the mitochondria and may influence cellular
sensitization to apoptotic stimuli via mitochondrial membrane
­destabilization29. Distinct from TMEM141 and TMEM14A,
TMEM14C belongs to an uncharacterized protein family
UPF0136_TM that is presumably involved in heme biosynthesis30.
Consistently with our earlier speculation6, we predict that IMPs
with tightly packed helices (TMEM14A and TMEM14C) may
have a structural or transport role in the membrane, whereas IMPs
with loosely packed helices (HIGD1A, HIGD1B and TMEM141)
could be involved in signal transduction across the membrane.
We believe that efficient production of hIMPs by cell-free
expression in combination with robust structural analysis by NMR
Articles
spectroscopy has broader applications. In addition to facilitating
3D structure determination of membrane proteins by solution
NMR spectroscopy, it will benefit functional and biochemical
characterization of hIMPs, including individual antibody generation against hIMPs for proteomic and cell biological studies.
Engineered hIMPs could also constitute a building block in biomaterial and nanoscience research.
Methods
Methods and any associated references are available in the online
version of the paper.
Accession codes. PDB: HIGD1A, 2LOM; HIGD1B, 2LON;
FAM14B, 2LOQ; TMEM141, 2LOR; TMEM14A, 2LOO and
2LOP; and TMEM14C, 2LOS.
npg
© 2012 Nature America, Inc. All rights reserved.
Note: Supplementary information is available in the online version of the paper.
Acknowledgments
We thank G. Louie for comments in preparation of the manuscript, A.S. Arseniev
for suggestions on the spin-labeling procedure and S. Maslennikov for writing the
atomDistancer program. C.K. thanks the Pioneer Foundation for a Pioneer Fund
Postdoctoral Scholar Award. This work has been partly supported by US National
Institutes of Health (S.C.: GM098630, GM095623; A.S. and U.P.: GM094662,
GM094625 FDP, and GM54762), Incheon Free Economic Zone and the World Class
University Program (Korea).
AUTHOR CONTRIBUTIONS
C.K., I.M., W.K., R.R. and S.C. designed experiments, C.K., E.J.C.C., L.E. and J.H.J.K.
cloned hIMP targets, performed cell-free expression, evaluated protein expression
levels and detergent solubilization; C.K. and E.J.C.C. created single cysteine
mutants for PRE experiments, prepared isotopically labeled NMR spectroscopy
samples and samples for PRE measurements. C.K. and I.M. recorded NMR spectra and
evaluated NMR spectral quality of tested hIMPs; C.K., I.M., M.B., C.E., N.V., E.J.C.C.
and K.B. collected and assigned NMR spectra and analyzed data; I.M., M.B. and C.E.
calculated the structures. C.K., E.J.C.C., B.B. and P.A.S. analyzed HIGD1A antibody
specificity by western blot and by immunostaining; U.P. and A.S. calculated
modeling leverage based on hIMP structures; C.K., I.M., W.K. and S.C. wrote the
manuscript. All authors discussed the results and commented on the manuscript.
COMPETING FINANCIAL INTERESTS
The authors declare no competing financial interests.
Published online at http://www.nature.com/doifinder/10.1038/nmeth.2033.
Reprints and permissions information is available online at http://www.nature.
com/reprints/index.html.
1. Klammt, C. et al. High level cell-free expression and specific labeling
of integral membrane proteins. Eur. J. Biochem. 271, 568–580 (2004).
2. Wuu, J.J. & Swartz, J.R. High yield cell-free production of integral membrane
proteins without refolding or detergents. Biochim. Biophys. Acta 1778,
1237–1250 (2008).
3. Katzen, F. et al. Insertion of membrane proteins into discoidal membranes
using a cell-free protein expression approach. J. Proteome Res. 7,
3535–3542 (2008).
4. Kalmbach, R. et al. Functional cell-free synthesis of a seven helix
membrane protein: in situ insertion of bacteriorhodopsin into liposomes.
J. Mol. Biol. 371, 639–648 (2007).
5. Klammt, C. et al. Evaluation of detergents for the soluble expression of alphahelical and beta-barrel-type integral membrane proteins by a preparative scale
individual cell-free expression system. FEBS J. 272, 6024–6038 (2005).
6. Maslennikov, I. et al. Membrane domain structures of three classes of
histidine kinase receptors by cell-free expression and rapid NMR analysis.
Proc. Natl. Acad. Sci. USA 107, 10902–10907 (2010).
7. Klammt, C. et al. Polymer-based cell-free expression of ligand-binding
family B G-protein coupled receptors without detergents. Protein Sci. 20,
1030–1041 (2011).
8. Junge, F. et al. Modulation of G-protein coupled receptor sample quality
by modified cell-free expression protocols: a case study of the human
endothelin A receptor. J. Struct. Biol. 172, 94–106 (2010).
9. Keller, T. et al. Cell free expression and functional reconstitution of
eukaryotic drug transporters. Biochemistry 47, 4552–4564 (2008).
10. Klammt, C. et al. Functional analysis of cell-free-produced human endothelin
B receptor reveals transmembrane segment 1 as an essential area for ET-1
binding and homodimer formation. FEBS J. 274, 3257–3269 (2007).
11. Ishihara, G. et al. Expression of G protein coupled receptors in a cell-free
translational system using detergents and thioredoxin-fusion vectors.
Protein Expr. Purif. 41, 27–37 (2005).
12. Pervushin, K., Riek, R., Wider, G. & Wuthrich, K. Attenuated T2 relaxation
by mutual cancellation of dipole-dipole coupling and chemical shift
anisotropy indicates an avenue to NMR structures of very large biological
macromolecules in solution. Proc. Natl. Acad. Sci. USA 94, 12366–12371
(1997).
13. Gautier, A. et al. Structure determination of the seven-helix
transmembrane receptor sensory rhodopsin II by solution NMR
spectroscopy. Nat. Struct. Mol. Biol. 17, 768–774 (2010).
14. Van Horn, W.D. et al. Solution nuclear magnetic resonance structure
of membrane-integral diacylglycerol kinase. Science 324, 1726–1729
(2009).
15. Wang, J., Pielak, R.M., McClintock, M.A. & Chou, J.J. Solution structure
and functional analysis of the influenza B proton channel. Nat. Struct.
Mol. Biol. 16, 1267–1271 (2009).
16. Hiller, S. et al. Solution structure of the integral human membrane protein
VDAC-1 in detergent micelles. Science 321, 1206–1210 (2008).
17. Bayrhuber, M. et al. Structure of the human voltage-dependent anion
channel. Proc. Natl. Acad. Sci. USA 105, 15370–15375 (2008).
18. Page, R.C. et al. Backbone structure of a small helical integral membrane
protein: A unique structural characterization. Protein Sci. 18, 134–146
(2009).
19. Liang, B., Bushweller, J.H. & Tamm, L.K. Site-directed parallel spinlabeling and paramagnetic relaxation enhancement in structure
determination of membrane proteins by solution NMR spectroscopy. J. Am.
Chem. Soc. 128, 4389–4397 (2006).
20. Iwahara, J., Schwieters, C.D. & Clore, G.M. Ensemble approach for NMR
structure refinement against (1)H paramagnetic relaxation enhancement
data arising from a flexible paramagnetic group attached to a
macromolecule. J. Am. Chem. Soc. 126, 5879–5896 (2004).
21. Battiste, J.L. & Wagner, G. Utilization of site-directed spin labeling and
high-resolution heteronuclear nuclear magnetic resonance for global fold
determination of large proteins with limited nuclear overhauser effect
data. Biochemistry 39, 5355–5365 (2000).
22. Kay, L.E. Solution NMR spectroscopy of supra-molecular systems,
why bother? A methyl-TROSY view. J. Magn. Reson. 210, 159–170
(2011).
23. Rual, J.F. et al. Human ORFeome version 1.1: a platform for reverse
proteomics. Genome Res. 14, 2128–2135 (2004).
24. Sobhanifar, S. et al. Structural investigation of the C-terminal catalytic
fragment of presenilin 1. Proc. Natl. Acad. Sci. USA 107, 9644–9649
(2010).
25. Zhou, Y. et al. NMR solution structure of the integral membrane enzyme
DsbB: functional insights into DsbB-catalyzed disulfide bond formation.
Mol. Cell 31, 896–908 (2008).
26. Bedo, G. et al. Characterization of hypoxia induced gene 1: expression
during rat central nervous system maturation and evidence of antisense
RNA expression. Int. J. Dev. Biol. 49, 431–436 (2005).
27. Kasper, L.H. & Brindle, P.K. Mammalian gene expression program
resiliency: the roles of multiple coactivator mechanisms in hypoxiaresponsive transcription. Cell Cycle 5, 142–146 (2006).
28. Jiang, Z., Gui, S. & Zhang, Y. Analysis of differential gene expression by
fiber-optic BeadArray and pathway in prolactinomas. Endocrine 38,
360–368 (2010).
29. Cheriyath, V., Leaman, D.W. & Borden, E.C. Emerging roles of FAM14
family members (G1P3/ISG 6–16 and ISG12/IFI27) in innate immunity
and cancer. J. Interferon Cytokine Res. 31, 173–181 (2011).
30. Nilsson, R. et al. Discovery of genes essential for heme biosynthesis through
large-scale gene expression analysis. Cell Metab. 10, 119–130 (2009).
31. Laskowski, R.A., MacArthur, M.W., Moss, D.S. & Thornton, J.M. PROCHECK:
A program to check the stereochemical quality of protein structures.
J. Appl. Cryst. 26, 283–291 (1993).
nature methods | VOL.9 NO.8 | AUGUST 2012 | 839
© 2012 Nature America, Inc. All rights reserved.
npg
ONLINE METHODS
Expression vector design. To enable cloning of hIMPs from a
library of 3,270 hIMPs in Gateway entry vectors23, the pIVEX2.3d
vector (Roche Applied Science) was Gateway-adapted and optimized. We designed two vectors, one containing several tags
that allow detection and purification and one without tag,
which was used for NMR spectroscopy sample preparation.
pIVEX 2.3d was supplemented with a 5′ attR1 site (5′-ACAA
GTTTGTACAAAAAAGCAGGCTTA-3′) and a 3′ attR2 site
(5′-GACCCAGCTTTCTTGTACAAAGTGGTT-3′), a thrombin
cleavage site (5′-GCTGCCACGCGGCACCAG-3′), a factor
Xa cleavage site (5′-ATCGAGGGCCGT-3′) and a StrepII tag
(5′-TGGAGCCACCCGCAGTTCGAAAAA-3′) using suitable
oligo­nucleotide primers with suitable restriction sites and standard
polymerase chain reaction techniques with Vent DNA polymerase
(New England Biolab (NEB)) following standard protocols for
Gateway destination vector creation (Invitrogen). The resulting
pIVEX2.3d-Gateway-tag vector (p23-GWT) encodes a protein
with an N-terminal Gateway sequence (MTSLYKKVG) and a
C-terminal tag (Y(or C)PTFLYKVVLVPRGSHMIEGRWSHPQ
FEKYRAPGGGSHHHHHH) (Fig. 1a). For Gateway cloning of
nontagged hIMPs for NMR spectral quality evaluation, a second
vector pIVEX2.3d-Gateway-NMR (p23-GWN) was derived from
p23-GWT by introducing a stop codon (TAA) after the 5′ att site,
resulting in translation of a short 9-amino-acid C-terminal attderived sequence (Y(or C)PTFLYKVV).
Cloning procedures. One hundred fifty hIMP targets in Gateway
entry vectors (Supplementary Table 6) were cloned into p23GWT and p23-GWN destination vectors using LR Clonase
(Invitrogen) according to the manufacturer’s protocol with optimizations. In particular, 1 µl entry clone (150 ng/µl), 1 µl destination vector (150 ng/µl), 2 µl 5× LR Clonase reaction buffer,
4 µl TE buffer (pH 8.0) and 2 µl Gateway LR Clonase II enzyme
mix (Invitrogen) were mixed and incubated for 60 min at 25 °C.
Subsequently, 1 µl proteinase K solution was added and LR reaction mixture was incubated for 10 min at 37 °C. One microliter
of LR reaction was transformed into 10 µl of DH5α chemically
competent cells (Invitrogen) and plated on LB plates containing
100 µg/ml ampicillin. Single colonies were picked and grown overnight in 5 ml TB medium with 0.1 mg/ml ampicillin and purified
using a Miniprep kit (Qiagen). The manufacturer’s protocol was
optimized to enhance plasmid yield and purity. In particular, cells
from 5-ml overnight cultures in TB medium were resuspended
in 250 µl buffer P1, and 350 µl buffer P2 was added and mixed
by gentle inversion. After 5 min, 450 µl of buffer N3 was added,
mixed and centrifuged for 10 min at 20,000g. The supernatant was
transferred to the QIAprep spin column connected to a vacuum
manifold (Qiagen). The column was washed with 500 µl buffer
PB and 750 µl buffer PE and subsequently centrifuged for 1 min
at 20,000g to remove residual ethanol. Plasmid DNA was eluted
with 75 µl buffer EB. Plasmids were checked by DNA sequencing
and used for cell-free expression.
Expression constructs for NMR spectroscopy structural studies of HIGD1A, HIGD1B, TMEM14A, FAM14B, TMEM141 and
TMEM14C were obtained by site-directed mutagenesis, introducing a stop codon (TAA) after the hIMP-encoding genes in
corresponding p23-GWN vectors. Cysteine residues in HIGD1A,
HIGD1B, TMEM14A, FAM14B, TMEM141 and TMEM14C as
nature methods
well as serine residues in HIGD1B, TMEM14A and TMEM141
for different cysteine constructs, were introduced by site-directed
mutagenesis. In particular, primers were designed as described
elsewhere32 and quick-change reactions were carried out using
1 µl HotStar polymerase (Qiagen), 1× HotStar buffer, 2% DMSO,
0.2 µM primers and 3–5 µg/ml template DNA in 50 µl of reaction
volume. PCR was set up in a thermocyler (Techne) at 95 °C for
0.5 min and cycled 18 times at 95 °C for 0.5 min, 55 °C for 100 s,
68 °C for 10 min with the final extension time of 30 min at 68 °C.
Parental DNA was digested with DpnI (NEB) by adding 1 µl
enzyme, incubated for 3 h at 37 °C and subsequently purified by
a Nucleotide purification kit (Qiagen) with elution in 30 µl H2O.
Seven microliters of DNA was used to transform 25 µl of DH5α
chemically competent cells (Invitrogen).
For transient expression of HIGD1A-GFP and HIGD1B-GFP
in HEK293T cells, the pTT5 expression vector33 was Gatewayadapted with suitable oligonucleotide primers as described above,
resulting in the pTT5-Gateway destination vector encoding an
N-terminal Human IgG κ signal peptide (METDTLLLWVLLLW
VPGSTGAGS) followed by a His9 tag and C-terminal translated
GFP. Sequences encoding HIGD1A and HIGD1B in Gatewayentry vectors were cloned into pTT5-Gateway as described above.
Plasmids were checked by DNA sequencing and amplified by
HighSpeed Plasmid Maxi Kit (Qiagen).
Cell-free expression. We established, optimized and fine-tuned
for expression of IMPs a preparative high-throughput E. coli–
based cell-free expression system. Chemicals for cell-free expression were purchased from Sigma-Aldrich. Stable isotope–labeled
amino acids and amino acid mixtures were purchased from
Cambridge Isotope Laboratories unless otherwise stated. hIMPs
were produced in an individual continuous exchange cell-free system according to previously described protocols with additional
optimization. In short, cell-free extracts were prepared from the
E. coli strain A19 as described previsously1,34, T7 RNA polymerase was expressed using the pT7-911Q plasmid35 and purified
as described previously36. Analytical-scale reactions were performed in 20 kDa molecular weight cutoff (MWCO) Mini SlideA-Lyzers (Thermo Scientific) using 70 µl of reaction mixture and
1 ml of feeding mixture. Mini Slide-A-Lyzers were placed in a
custom made 24-well plastic block holding the feeding mixture
and incubated in a shaker (New Brunswick Scientific) (Fig. 1a)
for ~15 h at 30 °C at 160 r.p.m. Preparative scale cell-free reactions were performed in 20-kDa MWCO Slide-A-Lyzers (Thermo
Scientific) using 2–4 ml of reaction mixture set with the 1:17 volume ratio between reaction and feeding mixture. Slide-A-Lyzers
were placed in a suitable plastic box holding the feeding mixture
and incubated in a shaker (New Brunswick Scientific) for ~15 h
at 30 °C at 140 r.p.m. The conditions for the cell-free reaction
were as follows: reaction mixture and feeding mixture contained
230 mM potassium acetate, 13 mM magnesium acetate, 100 mM
Hepes-KOH pH 8.0, 3.5 mM Tris-acetate pH 8.2, 0.2 mM folinic
acid, 0.05% sodium azide, 2% polyethyleneglycol 8000, 2 mM
Tris(2-carboxyethyl) phosphine hydrochloride (TCEP) (Thermo
Scientific), 1.2 mM ATP, 0.8 mM each of CTP, UTP and GTP,
20 mM acetyl phosphate (Fluka), 20 mM phosphoenol pyruvate (AppliChem), 1 tablet per 50 ml complete protease inhibitor (Roche Applied Science), 1 mM each amino acid, 40 µg/ml
pyruvate kinase (Roche Applied Science), 500 µg/ml E. coli tRNA
doi:10.1038/nmeth.2033
npg
© 2012 Nature America, Inc. All rights reserved.
mix (Roche Applied Science), 0.3 unit/µl RNase inhibitor
(SUPERase-In, Ambion), 0.5 unit/µl T7 RNA polymerase, 40% S30
extract and 10 µg/ml of pIVEX2.3d–derived plasmid DNA. For cellfree uniform 15N labeling, reaction mixture and feeding mixture
were supplemented with 0.5 mM of [15N]algal amino acid mixture
and 0.5 mM of 15N-labeled amino acids Asn, Cys, Gln and Trp. For
cell-free uniformly 15N-13C, 2H-15N, 2H-15N-13C labeling, reaction mixture and feeding mixture were supplemented with 0.5 mM
of correspondingly labeled amino acid mixtures. For combinatorial labeling of HIGD1A, HIGD1B and FAM14B combinations
of 15N-labeled Ala, Cys, Asp, Glu, Phe, Gly, Ile, Lys, Leu, Met,
Asn, Gln, Arg, Ser, Thr, Val, Trp and Tyr or 1-13C-labeled Ala,
Cys, Asp, Glu, Phe, Gly, Ile, Lys, Leu, Met, Pro, Gln, Ser, Val, Trp
and Tyr, and nonlabeled amino acids were used according to the
schemes given in Supplementary Table 1. Uniform 2H-15N and
2H-15N-13C labeling was efficiently done in H O.
2
Protein characterization. The Invitrogen gel ­electrophoresis ­system
was used for all SDS-gel analyses following the ­ manufacturer’s
protocol, using 12% NuPAGE Bis-Tris gels in 2-(N-morpholino)
ethanesulfonic acid (MES) buffer stained with Coomassie blue or
InstantBlue (Expedeon Protein Solutions).
The expression yield was quantified by estimating expression
based on Coomassie-stained protein band intensities for all 150
hIMPs. These intensities were compared to Coomassie-stained
standard bands of a known protein concentration. For western
blot analysis of cell-free expressed hIMPs, the gels were blotted
on a 0.45 µm Immobilon-P poly(vinylidene difluoride) membrane (Millipore) using Invitrogens Xcell IITM Blot Module for
1 h at 35 V. The membrane was then blocked for 1 h in blocking-buffer (1× PBS, 7% milk powder and 0.1% (w/v) Tween-20)
and subsequently incubated for 1 h with horseradish peroxidase
(HRP)-conjugated His6-tag antibody (ab1187, Abcam) using 1:
2,000 dilution in washing buffer (1× PBS and 0.1% Tween-20).
After extensive washing in washing buffer, the blots were analyzed by chemiluminescence (ECL western blot substrate, Thermo
Scientific) on X-ray film (CL-XPosure, Thermo Scientific) using
exposure times of 10–60 s.
For western blot analysis of HIGD1A and HIGD1A-GFP, we
used polyclonal anti-HIGD1A antibody raised in rabbit (Eton
Biosciences) from P-CF–expressed and LMPG-solubilized HIGD1A,
or rabbit anti-GFP (full-length) antibody (sc-8334, Santa Cruz
Biotechnology). The gels were blotted and blocked as described
above with subsequent 1 h incubation with anti-HIGD1A-IgG using
1:2,000 dilution or anti-GFP-IgG using 1:1,000 dilution in washing
buffer containing 7% milk powder. After incubation with primary
antibody, the membrane was washed 5 times with 100 ml washing buffer for 5 min each time, and subsequently incubated for 1 h
with secondary bovine anti-rabbit IgG-HRP (sc-2370, Santa Cruz
Biotechnology) using 1:3,000 dilution in washing buffer supplemented with 7% milk powder. After five 5-min washes with 100 ml
washing buffer, the blots were analyzed by chemiluminescence on
X-ray film using exposure of 0.5–5 min.
All cell-free expressed hIMPs were characterized by
SDS-PAGE (Supplementary Fig. 10). HIGD1A, HIGD1B,
TMEM14A, FAM14B, TMEM141 and TMEM14C were
analyzed by light scattering coupled with size-exclusion
chromatography and refracting index measurements (SECUV/LS/RI) (Supplementary Fig. 1). SEC-UV/LS/RI analysis of
doi:10.1038/nmeth.2033
hIMP-LMPG complexes was performed by measuring the relative
refractive index signal (Optilab rEX, Wyatt Technology), static
light scattering signals from three angles (45°, 90° and 135°) (miniDAWN TREOS, Wyatt Technology), and UV-light extinction at
280 nm (Waters 996 Photoiode Array Detector, Millipore) during
size-exclusion chromatography (HPLC, Waters 626 Pump, 600S
Controller, Millipore) with polymer column (Shodexfi Protein
KW-802.5). hIMPs were analyzed by injecting 100 µl of 200 µM
hIMP solubilized in LMPG into high-performance liquid chromatography (HPLC) buffer (20 mM MES-Bis-Tris pH 6.0 and 150 mM
NaCl) supplemented with 0.01% LMPG at 0.8 ml/min. The fractions,
containing target proteins, were concentrated in 5 kDa MWCO
Vivaspin2 concentrators (Sartorius Stedim Biotech) to 20–50 µl
and reloaded on the column. The oligomeric state of HIGD1B in
FC12, DDM and DM micelles was analyzed by SEC-UV/LS/RI.
HIGD1B was solubilized in 100 µl of a buffer (20 mM MES-Bis-Tris
pH 6.0 and 150 mM NaCl) containing selected detergent (20 mM,
30 mM and 75 mM for FC12, DDM and DM, respectively). The
final protein concentration was 150–200 µM. The samples were
injected into HPLC buffer supplemented with 1.6 mM FC12.
The data were collected and analyzed using the Astra V 5.3.2.12
Software (Wyatt Technology Corp.). The average molar weights of
the protein-detergent complex, the protein and the detergent fraction in the complex (Supplementary Fig. 1 and 2) were calculated
by the Protein Conjugate module of the Astra program. The oligomeric state of HIGD1B in DDM and DM could not be derived by
the Astra V Software from SEC-UV/LS/RI data due to the overlap
of protein-detergent micelles with empty detergent micelles as
shown for HIGD1B in the presence of DDM (Supplementary
Fig. 2c). Nevertheless the elution volumes of the HIGD1B protein
peak in the four detergents (Supplementary Fig. 2) are nearly
identical, which suggests that HIGD1B in the presence of DDM
and DM is most likely monomeric.
Detergent solubilization. All 150 P-CF-expressed hIMPs were
analyzed for detergent solubilization in seven different detergents.
Detergent solubilization was tested in 70 mM sodium dodecyl sulfate (SDS), 42 mM 1-myristoyl-2-hydroxy-sn-glycero-3-[phosphorac-(1-glycerol)] (LMPG), 100 mM n-decylphosphocholine
(FC10), 100 mM n-dodecylphosphocholine (FC12), 250 mM
n-decyl-β-d-maltoside (DM), mixture of 196 mM n-dodecyl-βd-maltoside (DDM) with 41 mM cholesteryl hemisuccinate (CHS)
and 100 mM lauryldimethylamine-N-oxide (LDAO). Seven samples of 7 µl of P-CF precipitate resuspended in buffer (20 mM
Tris pH 7.4 and 150 mM NaCl) were centrifuged for 10 min
at 20,000g. The supernatant was removed, and 7 µl of SDS, LMPG,
FC10, FC12, DM, DDM/CHS and LDAO were added to the
respective precipitate samples. The precipitate was resuspended
by pipetting 10–20 times and incubated for 1 h at 37 °C. The
residual precipitate was pelleted by centrifugation for 15 min
at 20,000g and 1–4 µl of the supernatant, depending on hIMP
expression, was loaded with 5 µl of 2× SDS sample buffer on a
12% NuPAGE Bis-Tris gel (Invitrogen) and run in MES buffer for
52 min at 200 V. The gel was stained with InstantBlue and analyzed for expression based on band intensity.
NMR spectroscopy sample preparation. All hIMPs were
expressed as precipitate (P-CF) in the absence of detergents1.
The only gene expressing in the cell-free system is the target
nature methods
© 2012 Nature America, Inc. All rights reserved.
npg
protein gene; therefore, the only labeled protein is the target protein. This, in combination with the fact that the SEC-UV/LS/RI data
showed our target proteins to be homogeneous protein-detergent
complexes, allows us to conclude that co-precipitated endogenous cell-free extract proteins will not influence NMR structural studies. Therefore, additional purification of the proteins
is not crucial. Precipitated recombinant proteins were removed
from the reaction mixture by centrifugation at 20,000g for 15 min
and washed in two steps. First, to remove co-precipitated RNA,
precipitates were suspended in 50% volume equal to the reaction mixture volume in 20 mM MES-Bis-Tris buffer pH 6.0, 0.01
mg/ml RNase A and shaken at 900 r.p.m. and 37 °C for 30 min.
After incubation, precipitates were collected by centrifugation at
20,000g for 10 min and suspended in 100% volume equal to the
reaction mixture volume in NMR buffer (20 mM MES-Bis-Tris
pH 6.0). NMR spectroscopy samples were prepared from washed
precipitate of 1–4 ml reaction mixture by solubilization in 300 µl
3% (wt/vol) LMPG in NMR buffer for all tested hIMPs except
HIGD1A, HIGD1B and TMEM141, which were solubilized in 2%
LMPG (wt/vol) in NMR buffer. The suspension was sonicated in a
water bath sonicator (Bransonic) for 1 min and subsequently incubated for 15 min with shaking at 900 r.p.m. and 37 °C, followed by
centrifugation at 20,000g for 10 min. NMR spectroscopy samples
were pH-adjusted and supplemented with 5% D2O and 0.5 mM
4,4-dimethyl-4-silapentane-1-sulfonic acid (DSS). For 13C and
15N NOESY NMR spectroscopy experiments requiring deuterated
detergent, hIMP NMR spectroscopy samples were prepared by
solubilization in 2% d27-LMPG (wt/vol) (FBReagents). D27-LMPG
was used to minimize the spectral distortion and the impact from
1H signals of the detergent in 13C-NMR spectroscopy experiments
with HIGD1B, TMEM14A and TMEM141. Shigemi NMR tubes
were used for solution NMR measurements. ‘Fingerprint’ spectra
of the cell-free-expressed hIMPs, categorized as good, are shown
in Figure 2 and Supplementary Figures 11 and 12.
Cysteine mutagenesis and spin-labeling. For PRE experiments, 5–9 single-cysteine mutants were prepared for HIGD1A,
HIGD1B, TMEM14A, FAM14B, TMEM141 and TMEM14C.
Cysteine-free mutants were prepared for HIGD1B, TMEM14A
and TMEM141 (Supplementary Table 3). The single cysteine
mutants prepared for each of the six selected hIMPs are listed
in Supplementary Table 3 and illustrated in Supplementary
Figure 4. Positions for cysteine introduction were chosen based
on the following criteria: mutations were located in regions containing helices, which were predicted by chemical shifts, and 3–4
residues adjoining the helix; each helix was labeled in at least two
positions close to its ends; mutated residues had minimal structural and/or functional importance; the preferred amino acids
for cysteine mutagenesis were serine, threonine and alanine; in
the case there were no appropriate serine, threonine or alanine;
residues close to the region of interest, valine, leucine, glutamine
and aromatic tyrosine and phenylalanine residues were the second
choice. Preservation of the structure in single-cysteine mutants
was tested by TROSY-HSQC experiments as disruption of helical
packing by the spin label would provide strong changes in chemical
shifts and would diminish the overall quality of the TROSY spectra. All single-cysteine mutants gave minimal changes in TROSYHSQC spectra upon introduction of the mutations. Even though
we followed the selection procedure described above, we still
nature methods
had few single-cysteine mutants for which satisfactory PRE data
were not obtained. For example, PRE effect in 1-oxyl-(2,2,5,5tetramethyl-Δ3-pyrroline-3-methyl)methanethiosulfonate
(MTSL)-labeled TMEM14A(S74C) affected only amide groups
from neighboring residues (not more than 10 amino acids
away from the labeled cysteine in the protein sequence). Other
than that, PRE data from the paramagnetic label attached to a
cysteine in non-transmembrane terminal helices require careful
analysis, because of high mobility of the helix and attached label.
The 15N-labeled single-cysteine mutants were prepared from
2–4 ml cell-free reaction mixture. To eliminate problems with
(i) incomplete reduction of the spin-label in a detergent-­solubilized
protein and (ii) changes in chemical shifts caused by incorporation of a label, we used parallel labeling with structurally similar paramagnetic and diamagnetic labels as suggested in ref. 19.
Every cysteine mutant was labeled with paramagnetic spin-label
MTSL and with diamagnetic label 1-acetyl-(2,2,5,5-tetrametyl
∆3-pyrroline-3-methyl)methanethiosulfonate (DML) (both
from Toronto Research Chemicals). The cysteine-free mutants
or cysteine-free wild-type proteins were used as a control of nonspecific MTSL binding to the protein and/or detergent and were
‘labeled’ with MTSL only. For the labeling, the 15N-labeled NMR
spectroscopy samples were split in half and supplemented with
5 mM MTSL or DML, solubilized in acetonitrile. After overnight
incubation at room temperature, the excess of MTSL and DML
was removed by washing in 5-kDa MWCO Vivaspin 2 concentrators (Sartorius Stedim Biotech). Labeled samples were washed
3 times by concentrating to 100 µl, resuspending in 2 ml NMR
buffer and concentrating to 100 µl. After the third wash the samples were concentrated to 300 µl, supplemented with 5% D2O and
0.5 mM DSS and measured in a Shigemi NMR tube.
NMR spectroscopy experiments. NMR spectra of hIMPs were
recorded at 37 °C on a Bruker AVANCE 700 MHz spectrometer
equipped with five radiofrequency channels and a triple-resonance
cryoprobe with a shielded z-gradient coil. For the combinatorial assignment, [15N,1H]TROSY-HSQC and 15N,1H plane of the
TROSY-HNCO12,37 were measured for each selectively 15N,13Clabeled sample. For the traditional assignment of backbone 1H,
15N and 13C resonances TROSY-based experiments, HNCA,
HNCO38, HNCACB, HNCOCA, HNCOCACB and HNCACO39
as well as gradient-enhanced 3D 1H-15N-NOESY-TROSY (mixing time, 120 ms) were used. Partial side chain assignment was
performed using 3D 1H-15N-NOESY-TROSY and 3D 1H-13CHSQC-NOESY–1H-15N-HSQC40 experiments. The PRE effect
was measured using [15N,1H]TROSY-HSQC spectra collected
for all cysteine mutants before spin labeling and after MTSL and
DML labeling. Protein localization within LMPG micelles was
checked by detection of a relaxation effect on [15N,1H]TROSYHSQC spectra of the hIMPs from water-soluble relaxation agent
Gd3+-DOTA (Molecular Probes)41. HIGD1B was measured with
different concentrations of Gd3+-DOTA (0 mM, 2.5 mM, 5.0 mM
and 10 mM). After analysis of the relaxation effect using ratios
of intensities in original (0 mM Gd3+-DOTA) and paramagnetic
samples (Supplementary Fig. 3a), the 5.0 mM concentration was
chosen to test the other hIMPs. The detected Gd3+-DOTA effect
(Supplementary Fig. 3) confirmed the transmembrane topologies for the calculated hIMPs structures. Protein-detergent NOEs
were derived using 3D 1H-15N-NOESY-TROSY experiments.
doi:10.1038/nmeth.2033
npg
© 2012 Nature America, Inc. All rights reserved.
The spectra were collected with 200-ms mixing time for the
15N,2H-labeled proteins (to achieve full protonation of amides,
the proteins were expressed and samples were prepared using
H2O) in protonated LMPG.
NMR signal assignment and spectra analysis. NMR spectra
were transformed using Topspin (Bruker Biospin) and ProSA
programs. Spectra analysis and assignment were performed using
the CARA program. Combinatorial assignment using CDL strategy6 was used to accelerate assignment procedure. Built on the
established principles of combinatorial assignment42,43, the CDL
strategy generates a sequence-dependent labeling scheme for 5–8
samples6. The samples are expressed in the P-CF mode and selectively 15N,13C-labeled according to this scheme (see, for example,
schemes for hIMPs in Supplementary Table 1). Analysis of the
cross-peaks in paired TROSY-HSQC and 2D HNCO experiments
allowed unambiguous assignment of those NH cross-peaks that
correspond to unique amino acid pairs in the protein sequence.
Other cross-peaks, with two or more possible assignments, are
assigned to an amino acid type. The schemes for CDL assignment were calculated using MCCL program6 and consisted of 6,
6 and 5 samples for HIGD1A, HIGD1B and FAM14B, respectively
(Supplementary Table 1). The selectivity of isotope labeling in
cell-free reaction can be affected by amino acid scrambling. The
most prominent examples are the pairs Gly-Ser, Asn-Gln, AspGlu and Asp-Asn. To avoid possible problems and facilitate CDL
assignment, we modified the MCCL algorithm (http://sbl.salk.
edu/combipro/) by incorporating user-defined identical labeling
for any selected group of amino acids. In case of any uncertainty,
the assignment can be verified using conventional sequential
assignment methods.
Calculation of PRE-based distance constraints. PRE distance
constraints were introduced for distances between an amide proton
and Cβ atom of residue, mutated to a cysteine for the paramagnetic labeling. Distance constraints were derived from the mea­
sured PRE effect using the procedures described in19,21,44,45. All
the spectra were transformed in the same way and the intensities
of 15N-1H cross-peaks in the MTSL (Ip) and DML (Id) samples
were measured using the CARA program. The ratios of intensities (Ip/Id) were normalized against a set of 8–12 highest Ip/Id
ratios, which were assumed to belong to cross-peaks unaffected
by PRE. For TMEM141, the PRE distance constraints were
derived by qualitative assessment of the Ip/Id ratios and were
categorized based on intra- or inter-helical contacts between the
spin label and the affected amide group as described45. PRE distance constraints for HIGD1A, HIGD1B, TMEM14A, FAM14B
and TMEM14C were calculated using the modified SolomonBloembergen equation (equation (5) in ref. 21). The transverse
relaxation rate enhancement was obtained from normalized
intensity ratios (Ip/Id) as previously described, and the correlation time for the electron-nuclear spin interaction was estimated
as the global rotational correlation time of the protein-detergent
complex calculated using the Stokes-Einstein equation (Table 1).
Such rough estimation of the correlation time is sufficient enough
because, as mentioned earlier21, even moderate 20% error in estimation of the correlation time gives only ~0.5 Å error in calculated distance. For cross-peaks with the ratios below 0.15, no
lower distance constraints were used, whereas upper constraints
doi:10.1038/nmeth.2033
were set to 12 Å for distances between stable regions and to 15 Å
if a spin label or amide group were located in flexible regions. For
cross-peaks with the ratios above 0.9, only lower distance constraints equal to 25 Å were introduced. The upper distance constrains between flexible or unstructured regions were excluded
from calculation. The upper and lower distance constraints for the
peaks with Ip/Id ratios between 0.15 and 0.9 were generated from
PRE-calculated distances using ± 4 Å margins. During structure
calculation the margins were reduced to ± 3 Å for constraints
between structured transmembrane regions if this change did not
increase the Cyana penalty function. We assumed that the initial
± 4 Å margins in distance constraints are sufficient to cover the
possible errors resulting from the use of a uniform correlation
time, the uncertainty of the estimation of the intrinsic relaxation
rates and the ‘r−6’-averaging of the nitroxide group motion. It was
shown that the whole side chain of cysteine with attached MTSL
(R1) has reduced mobility in both water-soluble and membrane
proteins46,47. In the studied membrane leucine transporter46, the
nitroxide rings interact with the hydrophobic protein surface,
thus fixing the whole R1 side chain.
We found that PRE-derived lower distance constraints are
important for the calculation of helical bundle structures and
especially for the determination of relative orientations of the
transmembrane helices. As the only physical constraints of an
α-helical bundle are van der Waals interactions, it has to have
additional lower limit constrains from experimental data to prevent collapsing of the bundle. In contrast, others had reported
that lower constraints were insignificant in the calculation of the
β-barrel structure of OmpA19. This is not surprising because a
β-barrel is restrained by the local geometry of inter-strand contacts and by the constant network of these contacts.
As precision of the PRE-derived distance constraints is low,
for successful structure calculation it is important to obtain as
many meaningful constraints as possible. In theory, for every
HN group from a given helical region the number of PRE-based
distance constraints should be equal to the number of cysteine
mutants used for spin-labeling. In reality, this number is lower
due to signal overlapping and meaningless constraints like those
between neighboring atoms within the same transmembrane
helices. The average number of upper distance constraints per
restrained residue ranged from 3.63 (HIGD1A) to 4.64 (FAM14B)
for the studied hIMPs. According to ref. 48, approximate global
fold can be determined with as few as 1.4 constraints per residue,
whereas at least 3 constraints per residue are required for low- to
medium-resolution NMR spectroscopy structures.
Structure calculation and analysis. The 13Cα, 13Cβ and 13CO
chemical shift deviations from random coil values were used to
define backbone torsion angle restraints 49. Sequential distance
constraints were derived from the integral intensities of NOE
cross-peaks measured in 3D 15N-resolved TROSY-[1H;1H]NOESY (mixing time 120 ms). The hydrogen bond constraints
were generated for the helical regions defined by chemical shift
analysis. An interactive procedure, which included structure
calculation by the CYANA program50 followed by the distance
constraints refinement, was used to calculate the backbone spatial
structures of the hIMPs. The structures were calculated using a
simulated annealing protocol (1,000 high-temperature steps followed by 9,000–11,000 cooling down steps and 1,300 steps of a
nature methods
npg
© 2012 Nature America, Inc. All rights reserved.
conjugate gradient minimization) and the default CYANA force
field. The summary of the constraints used in the calculation of
the structures is presented in Table 1. Long-distance constraints
used in structure calculation are shown in Figure 3b. The 20 conformers with the lowest target function of the last CYANA calculation cycle were selected from 200 calculated structures. The
helical packing parameters, such as interhelical crossing angles
and helical kinks, were derived for the final sets of 20 structures
with the Helix Packing Pair51 and Molmol52 programs. The structures were visualized and analyzed in Molmol program; statistics
for interatomic distances (average value and deviation) in the sets
of structures were calculated using atomDistancer program.
Cell culture. Human embryonic kidney cells HEK293T were
grown in DMEM medium (Mediatech) supplemented with 10%
FCS in humidified incubator at 37 °C with 5% CO2. Primary
cultures of hippocampal neurons were prepared from 0–2-d-old
Sprague Dawley rat pups using a modification of a previously
described method53. Briefly, the hippocampi were dissected from
brain and dissociated with papain (Worthington), and the neurons were plated at 25,000 cells/cm2 onto 12-mm glass cover slips
(Warner Instruments) coated with 0.2 mg/ml poly-d-lysine (BD).
Hippocampal neurons were cultured in Neurobasal medium supplement with B27, 100 U/ml streptomycin and 100 µg/ml penicillin (Invitrogen) at 37 °C and in 5% CO2 for 10–14 d. The medium
was replaced the day after plating and twice weekly thereafter. All
the procedures were approved by the Salk Institute’s Institutional
Animal Care and Use Committee.
Heterologous expression of hIMP proteins in human cell line.
HEK293T cells were grown to 90–95% confluence and transiently transfected with DNA encoding for the indicated proteins using Lipofectamine 2000 (Invitrogen) according to the
manufacturer’s instructions. Forty-eight hours after transfection,
cell culture medium was aspirated, and cells were washed twice
with PBS. Cell were then placed on ice and lysed with mild lysis
solution (Immunocatcher kit, CytoSignal Research Products)
supplemented with protein inhibitors (Complete, Mini, Roche
Applied Science).
Immunofluorescence and imaging. Culture medium was aspirated from cell cultures, followed by a brief wash with PBS. Cells
were then fixed with 4% paraformaldehyde solution in PBS for
10 min at room temperature, washed and permeabilized with 0.1%
Triton-X in PBS (10 min). Unspecific binding was blocked by
30 min incubation with 3% BSA solution in PBS. Cells were then
incubated with primary rabbit anti-HIGD1A antibody diluted
1:5,000 in blocking solution for 1 h, washed thoroughly with PBS
and incubated over 1 h with secondary anti-rabbit Alexa Fluor
647–conjugated antibody (1:400 dilution, A-21244, Invitrogen).
After the final wash with PBS, cells were mounted using ProLong
Gold antifade reagent (Invitrogen) and imaged with a laser scanning confocal microscope (Zeiss LSM 710) using a 63× objective
with oil immersion. The differential image contrast (DIC) was
corrected using ImageJ Pseudoflatfield Filter followed by adjustment of brightness and contrast.
Pictures of live HEK293T cell transfected with vectors encoding GFP fusion proteins were taken using a Nikon Eclipse TE300
inverted microscope equipped with a high-pressure mercury
nature methods
lamp and a GFP set of filters and beam splitters (Excitation filter
HQ480/40x; dichroic mirror Q505LP, Emission filter HQ525/
50m, Chroma) and a Nikon D70 digital camera.
Modeling leverage calculations. The modeling leverage of
the six hIMP NMR spectroscopy structures was estimated
by the ModPipe, comparative modeling pipeline54 accessible
through the ModWeb web server (http://salilab.org/modweb/)
(Supplementary Table 5). We relied on the ModWeb option that
accepts a protein structure as input, calculates a multiple sequence
profile and identifies all homologous sequences in the UniProtKB
database55, followed by modeling these homologs based on the
user-provided structure. These models are available in ModBase
through a summary page (http://modbase.compbio.ucsf.edu/
modbase-cgi/model_leverage.cgi?type=master_salk). On average,
each structure allowed us to model with relatively high accuracy
171 related unique protein sequences, based on more than 30%
sequence identity and using at least 50% of the residues in the
structures as templates. Another 114 protein sequences on average
could also be modeled, but at lower accuracy, primarily because
of the target-template alignment errors.
32. Zheng, L., Baumann, U. & Reymond, J.L. An efficient one-step sitedirected and site-saturation mutagenesis protocol. Nucleic Acids Res. 32,
e115 (2004).
33. Shi, C. et al. Purification and characterization of a recombinant G-proteincoupled receptor, Saccharomyces cerevisiae Ste2p, transiently expressed in
HEK293 EBNA1 cells. Biochemistry 44, 15705–15714 (2005).
34. Klammt, C., Schwarz, D., Dotsch, V. & Bernhard, F. Cell-free production of
integral membrane proteins on a preparative scale. Methods Mol. Biol. 375,
57–78 (2007).
35. Ichetovkin, I.E., Abramochkin, G. & Shrader, T.E. Substrate recognition by
the leucyl/phenylalanyl-tRNA-protein transferase. Conservation within the
enzyme family and localization to the trypsin-resistant domain. J. Biol.
Chem. 272, 33009–33014 (1997).
36. Savage, D.F. et al. Cell-free complements in vivo expression of the E. coli
membrane proteome. Protein Sci. 16, 966–976 (2007).
37. Riek, R. et al. Solution NMR techniques for large molecular and
supramolecular structures. J. Am. Chem. Soc. 124, 12144–12153 (2002).
38. Salzmann, M. et al. TROSY in triple-resonance experiments: new
perspectives for sequential NMR assignment of large proteins. Proc. Natl.
Acad. Sci. USA 95, 13585–13590 (1998).
39. Salzmann, M. et al. TROSY-type triple-resonance experiments for sequential
NMR assignments of large proteins. J. Am. Chem. Soc. 121, 844–848 (1999).
40. Diercks, T., Coles, M. & Kessler, H. An efficient strategy for assignment of
cross-peaks in 3D heteronuclear NOESY experiments. J. Biomol. NMR 15,
177–180 (1999).
41. Hilty, C., Wider, G., Fernandez, C. & Wuthrich, K. Membrane protein-lipid
interactions in mixed micelles studied by NMR spectroscopy with the use
of paramagnetic reagents. ChemBioChem 5, 467–473 (2004).
42. Yabuki, T. et al. Dual amino acid-selective and site-directed stable-isotope
labeling of the human c-Ha-Ras protein by cell-free synthesis. J. Biomol.
NMR 11, 295–306 (1998).
43. Kainosho, M. & Tsuji, T. Assignment of the three methionyl carbonyl
carbon resonances in Streptomyces subtilisin inhibitor by a carbon-13 and
nitrogen-15 double-labeling technique. A new strategy for structural
studies of proteins in solution. Biochemistry 21, 6273–6279 (1982).
44. Van Horn, W.D., Beel, A.J., Kang, C. & Sanders, C.R. The impact of
window functions on NMR-based paramagnetic relaxation enhancement
measurements in membrane proteins. Biochim. Biophys. Acta 1798,
140–149 (2010).
45. Roosild, T.P. et al. NMR structure of Mistic, a membrane-integrating
protein for membrane protein expression. Science 307, 1317–1321 (2005).
46. Kroncke, B.M., Horanyi, P.S. & Columbus, L. Structural origins of nitroxide
side chain dynamics on membrane protein alpha-helical sites. Biochemistry
49, 10045–10060 (2010).
47. Langen, R., Oh, K.J., Cascio, D. & Hubbell, W.L. Crystal structures of spin
labeled T4 lysozyme mutants: implications for the interpretation of EPR
spectra in terms of structure. Biochemistry 39, 8396–8405 (2000).
doi:10.1038/nmeth.2033
52. Koradi, R., Billeter, M. & Wuthrich, K. MOLMOL: a program for display
and analysis of macromolecular structures. J. Mol. Graph. 14, 51–55
(1996).
53. Brewer, G.J., Torricelli, J.R., Evege, E.K. & Price, P.J. Optimized survival of
hippocampal neurons in B27-supplemented Neurobasal, a new serum-free
medium combination. J. Neurosci. Res. 35, 567–576 (1993).
54. Pieper, U. et al. ModBase, a database of annotated comparative protein
structure models, and associated resources. Nucleic Acids Res. 39,
D465–D474 (2011).
55. Bairoch, A. et al. The Universal Protein Resource (UniProt). Nucleic Acids
Res. 33, D154–D159 (2005).
npg
© 2012 Nature America, Inc. All rights reserved.
48. Clore, G.M., Robien, M.A. & Gronenborn, A.M. Exploring the limits of
precision and accuracy of protein structures determined by nuclear
magnetic resonance spectroscopy. J. Mol. Biol. 231, 82–102 (1993).
49. Lubingbuhl, P., Szyperski, T. & Wuthrich, K. Statistical basis for the use of
13Ca chemical shifts in protein structure determination. J. Magn. Reson.
B. 109, 229–233 (1995).
50. Guntert, P. Automated NMR structure calculation with CYANA. Methods
Mol. Biol. 278, 353–378 (2004).
51. Dalton, J.A., Michalopoulos, I. & Westhead, D.R. Calculation of helix
packing angles in protein structures. Bioinformatics 19, 1298–1299
(2003).
doi:10.1038/nmeth.2033
nature methods